• Aucun résultat trouvé

STANDARD SURFACES AND NODAL CURVES IN SYMPLECTIC 4-MANIFOLDS Michael Usher Abstract

N/A
N/A
Protected

Academic year: 2022

Partager "STANDARD SURFACES AND NODAL CURVES IN SYMPLECTIC 4-MANIFOLDS Michael Usher Abstract"

Copied!
54
0
0

Texte intégral

(1)

77(2007) 237-290

STANDARD SURFACES AND NODAL CURVES IN SYMPLECTIC 4-MANIFOLDS

Michael Usher

Abstract

Continuing the program of [DS] and [U1], we introduce re- finements of the Donaldson-Smith standard surface count which are designed to count nodal pseudoholomorphic curves and curves with a prescribed decomposition into reducible components. In cases where a corresponding analogue of the Gromov-Taubes in- variant is easy to define, our invariants agree with those analogues.

We also prove a vanishing result for some of the invariants that count nodal curves.

1. Introduction

Let (X, ω) be a closed symplectic 4-manifold. We assume that [ω]∈ H2(X,Z); however, the main theorems in this paper concern Gromov invariants, which are unchanged under deformations of the symplectic form, so since any symplectic form is deformation equivalent to an in- tegral form there is no real loss of generality here. According to [Do], if k is large enough, taking a suitable pair of sections of a line bundle L⊗k where L has Chern class [ω] and blowing X up at the common vanishing locus of these sections to obtain the new manifold X gives rise to a symplectic Lefschetz fibration f: X → CP1 (the exceptional curves of the blowup π: X → X appear as sections of f, while at other points x ∈ X, f(x) ∈ C∪ {∞} is the ratio of the two chosen sections of L⊗k at π(x) ∈ X). In other words, f is a fibration by Riemann surfaces over the complement of a finite set of critical values in S2, while near its critical pointsf is given in smooth local complex coordinates by f(z, w) = zw. Results of [Sm1] show that the critical points of f may be assumed to lie in separate fibers, and all fibers of f may be assumed irreducible. Once we choose a metric on X, Donald- son’s construction thus presents a suitable blowup of X as a smoothly CP1-parametrized family of Riemann surfaces, all but finitely of which are smooth and all of which are irreducible with at worst one ordinary double point. Where κX = c1(TX) is the canonical class of X, note

Received 04/15/2005.

237

(2)

that the adjunction formula gives the arithmetic genus of the fibers as g= 1 + (k2[ω]2+kκX ·ω)/2.

Beginning with the work of S. Donaldson and I. Smith in [DS], some efforts have recently been made toward determining whether such a Lefschetz fibration can shed light on any questions concerning pseudo- holomorphic curves in X. More specifically, for any natural number r Donaldson and Smith construct the relative Hilbert scheme, which is a smooth symplectic manifoldXr(f) with a mapF: Xr(f)→CP1whose fiber over a regular value tof f is the symmetric product Srf−1(t). If we choose an almost complex structure j on X with respect to which f is a pseudoholomorphic map, a j-holomorphic curve C in X which contains no fiber components will, by the positivity of intersections be- tweenj-holomorphic curves, meet each fiber inr := [C]·[f iber] points, counted with multiplicities. In other words, C∩f−1(t)∈Srf−1(t), so that, letting tvary,C gives rise to a sectionsC ofXr(f). Conversely, a section sof Xr(f) gives rise to a subset Cs of X (namely the union of all the points appearing in the divisorss(t) as tvaries), and fromj one may construct a (nongeneric and generally not even C1) almost com- plex structureJj with the property that C is a (possibly disconnected) j-holomorphic curve in X if and only ifsC is aJj-holomorphic section of Xr(f).

Accordingly, it seems reasonable to study pseudoholomorphic curves in X by studying pseudoholomorphic sections of Xr(f). If α ∈ H2(X;Z), thestandard surface count DSf(α) is defined in [Sm2] (and earlier in [DS] for α = κX) as the Gromov-Witten invariant which countsJ-holomorphic sectionsswhose corresponding setsCs are Poin- car´e dual to the classαand pass through a generic set ofd(α) = 122− κX ·α) points of X, where J is a generic almost complex structure on Xr(f). Smith shows in [Sm2] that there is at most one homotopy class cα of sections ssuch that Cs is Poincar´e dual toα, and moreover that the complex dimension of the space of J-holomorphic sections in this homotopy class is, for generic J, the aforementioned d(α), which the reader may recognize as the same as the expected dimension of j- holomorphic submanifolds ofXPoincar´e dual toα. Further, the moduli space ofJ-holomorphic sections in the homotopy classcαis compact for genericJifkis taken large enough. The moduli space in the definition of DSf is therefore a finite set, andDSf simply counts the members of this set with sign according to the usual (spectral-flow-based) prescription.

Donaldson and Smith have proven various results aboutDS, perhaps the most notable of which is the main theorem of [Sm2], which asserts that if α ∈H2(X;Z), if b+(X)> b1(X) + 1, and if the degreek of the Lefschetz fibration is high enough, then

(1.1) DSfα) =±DSfX −α)).

(3)

Their work has led to new, more symplectic proofs of various results in 4-dimensional symplectic topology which had previously been accessi- ble only by Seiberg-Witten theory (as an example we mention the main theorem of [DS], according to which X admits a symplectic surface Poincar´e dual to κX, again assuming b+(X) > b1(X) + 1). In [U1] it was shown that the invariantDSf agrees with the Gromov invariantGr which was introduced by C. Taubes in [T2] and which counts possibly- disconnected pseudoholomorphic submanifolds ofX Poincar´e dual to a given cohomology class. This in particular shows that DSf is indepen- dent of the choice of Lefschetz fibration structure, and, in combination with Smith’s duality theorem (1.1) and the fact that under a blowup π one hasGr(πα) =Gr(α), yields a new proof of the relation

Gr(α) =±Gr(κX −α)

ifb+(X)> b1(X) + 1, a result which had previously only been known as a shadow of the charge conjugation symmetry in Seiberg-Witten theory.

The information contained in the Gromov invariants comprises only a part of the data that might be extracted from pseudoholomorphic curves inX. The present paper aims to show that many of these addi- tional data can also be captured by Donaldson-Smith-type invariants.

For instance, Gr(α) counts all of the curves with any decomposition into connected components whose homology classes add up (counted with multiplicities) to α. It is natural to wish to keep track of the de- compositions of our curves into reducible components; accordingly we make the following:

Definition 1.1. Letα∈H2(X;Z). Let

α=β1+· · ·+βm+c1τ1+· · ·+cnτn

be a decomposition of α into distinct summands, where none of theβi satisfies βi2 = κX ·βi = 0, while the τi are distinct classes which are primitive in the latticeH2(X;Z) and all satisfyτi2X ·τi= 0. Then

Gr(α;β1, . . . , βm, c1τ1,· · · , cnτn)

is the invariant counting ordered (m+n)-tuples (C1, . . . , Cm+n) of trans- versely intersecting smooth pseudoholomorphic curves inX, where

(i) for 1≤i≤m,Ci is a connected curve Poincar´e dual to βi which passes through some prescribed generic set of d(βi) points;

(ii) form+ 1≤k≤m+n,Ckis a union of connected curves Poincar´e dual to classes lk,1τk,· · · , lk,pτk decorated with positive integer multiplicities mk,q with the property that

X

q

mk,qlk,q =ck.

(4)

The weight of each component of each such curve is to be determined according to the prescription given in the definition of the Gromov in- variant in [T2] (in particular, the components Ck,q in class lk,qτk are given the weight r(Ck,q, mk,q) specified in Section 3 of [T2]), and the contribution of the entire curve is the product of the weights of its com- ponents.

The objects counted byGr(α;α1, ..., αn) will then be reducible curves with smooth irreducible components and a total ofP

αi·αjnodes arising from intersections between these components. Gr(α) is the sum over all decompositions of α into classes which are pairwise orthogonal under the cup product of the

d(α)!

Q(d(αi)!)Gr(α;α1, . . . , αn);

in turn, one has

Gr(α;α1, . . . , αn) = Yn i=1

Gr(αii).

In Section 2, given a symplectic Lefschetz fibration f: X → S2 with sufficiently large fibers, by counting sections of a relative Hilbert scheme we construct a corresponding invariant gDSf(α;α1, . . . , αn) pro- vided that none of the αi can be written as mβ where m >1 and β is Poincar´e dual to either a symplectic square-zero torus or a symplectic (−1)-sphere. Further:

Theorem 1.2. (QP(d(αd(αi))!

i)!)Gr(α;α1, . . . , αn) =DSgf(α;α1, . . . , αn)pro- vided that the degree of the fibration is large enough that h[ωX],[Φ]i >

X]·α.

The sections s counted by gDSf(α;α1, . . . , αn) correspond tautologi- cally to curves Cs =∪Csi inX with each Csi Poincar´e dual to αi. The Csi will be symplectic, and Proposition 2.5 guarantees that they will intersect each other positively, so there will exist an almost complex structure making Cs holomorphic. However, ifs1 and s2 are two differ- ent sections in the moduli space enumerated by DSgf(α;α1, . . . , αn), it is unclear whether there will exist a single almost complex structure on X making both Cs1 and Cs2 holomorphic.

The almost complex structures on Xr(f) used in the definition of gDS are, quite crucially, required to preserve the tangent space to the diagonal stratum consisting of divisors with one or more points re- peated. One might hope to define analogous invariants which agree with Gr(α;α1, . . . , αn) using arbitrary almost complex structures on Xr(f). If one could do this, though, the arguments reviewed in Section 4 would rather quickly enable one to conclude thatGr(α;α1, . . . , αn) = 0

(5)

whenever α has larger pairing with the symplectic form than does the canonical class and αi·αj = 0 for i6=j. However, this is not the case:

the manifold considered in [MT] admits a symplectic form such that, for certain primitive, orthogonal, square-zero classesα,β,γ, andδeach with positive symplectic area, the canonical class is 2(α+β +γ) but the invariant Gr(2(α+β+γ) +δ;α, β, γ, α+β+γ+δ) is nonzero.

While the Gromov–Taubes invariant restricts attention to curves whose components are all covers of embedded curves which do not in- tersect each other, it is natural to hope for information about curves Poincar´e dual toαhaving some numbernof transverse self-intersections.

One might like to define an analogueGrn(α) of the Gromov–Taubes in- variant counting such curves, but as we review in Section 3, owing to issues relating to multiple covers it is somewhat unclear what the def- inition of such an invariant should be in general. If one imposes some rather stringent conditions on α (α should be “n-semisimple” in the sense of Definition 3.1), there is however a natural such choice.

Note that for arbitrary α and n, following [RT] one may define an invariantRTn(α) which might naively be viewed as a count ofconnected pseudoholomorphic curves Poincar´e dual to α with n self-intersections by enumerating solutionsu: Σg→Xof the equation ( ¯∂ju) =ν(x, u(x)) for generic j and “inhomogeneous term” ν, where the genus g of the source curve is given in accordance with the adjunction formula by 2g− 2 = α2X ·α −2n. (Note that the nontrivial dependence of ν on x prevents multiple cover problems from arising.) In the case n = 0, the main theorem of [IP1] provides a universal formula equatingGr(α) with a certain combination of the Ruan–Tian invariantsRT0. The proof of that theorem goes through easily to show that in the case when α is n-semisimple, there exists a similar formula equating Grn(α) with a combination of Ruan–Tian invariants. We mention also that, again as an artifact of the multiple cover problem, the Ruan–Tian invariants are obliged to take values inQrather thanZ. Gr(α), on the other hand, is an integer-valued invariant.

By combining the approaches of [DS] and [L1], in the presence of a Lefschetz fibration f: X → S2 we construct in Section 3 an integer- valued invariantFDSnf(α−2P

ei) which we conjecture to be an appro- priate candidate for a “nodal version” Grn(α) of the Gromov invariant for general classesα(after perhaps dividing by n! to account for a sym- metry in the construction). Pleasingly, the technical difficulties that often arise in defining invariants like Grn(α) do not affect FDS: since FDS counts sections of a (singular) fibration, which of course neces- sarily represent a primitive homology class in the total space, we need not worry about multiple covers; further, the fact that any bubbles that form in the limit of a sequence of holomorphic sections must be con- tained in the fibers of the fibration turns out (via an easy elaboration

(6)

of a dimension computation from [DS]) to generically rule out bubbling as well. In principle, though, FDSnf might depend on the choice of Lefschetz fibration f.

Note that ifπ: X →Xis a blowup with exceptional divisor Poincar´e dual to ǫ, whenever Grn(β) is defined we will have (Grn)X(β+ǫ) = (Grn)X(β) (here and elsewhere we use the same notation for β ∈ H2(X;Z) and πβ ∈ H2(X;Z)), as the curves contributing to (Grn)X(β) generically miss the point being blown up, and so the unions of their proper transforms with the exceptional divisor will be precisely the curves contributing to (Grn)X(β+ǫ). With this said, we formulate:

Conjecture 1.3. Let (X, ω) be a symplectic 4-manifold and α ∈ H2(X;Z), andf: X →S2 a Lefschetz fibration obtained from a suffi- ciently high-degree Lefschetz pencil onX, with the exceptional divisors of the blowup X → X Poincar´e dual to the classes ǫ1, . . . , ǫN. Then the family Donaldson–Smith invariants

FDSnf Ã

α+ XN

i=1

ǫi−2 Xn k=1

ek

!

are independent of the choice of f, and have a general expression in terms of the Ruan–Tian invariants of X.

Note that this conjectural general expression would then produce an integer by taking appropriate combinations of the (a priori only rational) Ruan–Tian invariants, similarly to the formula of [IP1].

In light of the behavior ofGrnunder blowups, Theorem 3.8 amounts to the statement that:

Theorem 1.4. If α is strongly n-semisimple, then Conjecture 1.3 holds for α; more specifically, we have

FDSnf Ã

α+ XN

i=1

ǫi−2 Xn k=1

ek

!

=n!Grn(α).

We also prove that FDS vanishes under certain circumstances. This result depends heavily on the constructions used by Smith in [Sm2]

to prove his duality theorem, and so we review these constructions in Section 4. Section 5 is then devoted to a proof of the following theorem.

Theorem 1.5. Ifb+(X)> b1(X) + 4n+ 1, then for allα ∈H2(X;Z) such that r = hα,[Φ]i satisfies r > max{g(Φ) + 3n+d(α),(4g(Φ)− 11)/3}, either FDSnf(α−2P

ei) = 0 or there exists an almost complex structure j on X compatible with the fibration f: X → S2 which si- multaneously admits holomorphic curves C and DPoincar´e dual to the classes α and κX −α. In particular, FDSnf(α−2P

ei) = 0 if α has larger pairing with the symplectic form than does κX.

(7)

Note that in the Lefschetz fibrations obtained from degree-kLefschetz pencils on some fixed symplectic manifold (X, ω), the number N of exceptional sections is k2[ω]2 while the number 2g(Φ)−2 is asymptotic tok2[ω]2, so the invariants

FDSnf Ã

α+ XN

i=1

ǫi−2X

k

ek

!

considered in Conjecture 1.3 all eventually satisfy the restriction on r in Theorem 1.5.

The almost complex structure in the second alternative in Theorem 1.5 cannot be taken to be regular (in the sense that the moduli spaces MjX(β) ofj-holomorphic curves Poincar´e dual toβ are of the expected dimension); the most we can say appears to be that it can be taken to be a member of a regular 4n-real-dimensional family of almost complex structures,i.e., a family of almost complex structures{jb}parametrized by elements b of an open set in R4n such that the spaces {(b, C)|C ∈ MjXb(β)}are of the expected real dimension 2d(β) + 4nnear each (b, C) such that C has no multiply-covered components. Also, if X is in fact K¨ahler and admits a compatible integrable complex structure j0 with respect to which the fibration f is holomorphic, then we can take thej in Theorem 1.5 equal to j0.

In fact, if we could take j to be regular, then we could rule out the second alternative in Theorem 1.5 entirely (when n >0) using the following argument: the invariant vanishes trivially when d(α) < n, so we can assume d(α) = −12α ·(κ−α) > 0. But then our curves Poincar´e dual to αand κ−αhave negative intersection number, which is only possible if they share one or more components of negative square.

For generic j, a virtual dimension computation shows that the only irreducible j-holomorphic curves of negative square are (−1)-spheres.

Moreover whatever (−1)-spheres appear in X must be disjoint, since if they were not, blowing one of two intersecting (−1)-spheres down would cause the image of the other to be a symplectic sphere of nonnegative self-intersection, which (by a result of [McD]) would force X to have b+ = 1, which we assumed it did not. Ignoring all the (−1)-spheres in C and D and taking the union of what is left over would then give a j-holomorphic curve Poincar´e dual to a class κX −P

aiei where the ei are classes of (−1)-spheres with ei·ek = 0 for i 6= k and where at least oneai≥2. But one easily finds d(κX −P

aiei)<0, so this too is impossible for generic j. For nongeneric j, this argument breaks down because of the possibility that C and D might share components of negative square and negative expected dimension, and there is a wider diversity of possible homology classes of such curves.

The final section of the paper contains proofs of two technical results that are used in the proofs of the main theorems. First, we show that

(8)

the operation of blowing up a point can be performed in the almost complex category, a fact which does not seem to appear in the literature and whose proof is perhaps more subtle than one might anticipate. The paper then closes with a proof of the following result, which is necessary for the compactness argument that we use to justify the definition of our invariant FDS:

Theorem 1.6. Let F: Hr → D2 denote the r-fold relative Hilbert scheme of the map(z, w)7→zw,φ0the partial resolution mapF−1(0)→ Symr{zw = 0}, and ∆ ⊂ Hr the diagonal stratum. At any point p ∈

∆∩F−1(0) with φ0(p) ={(0,0), . . . ,(0,0)}, where Tp∆ is the tangent cone to ∆ at p, we have Tp∆⊂TpF−1(0).

We end the introduction with some remarks on the possible relation of FDS to (family) Seiberg–Witten theory. In [Sa] it was shown that where X is the product of R and a fibered three-manifold, so that X fibers over a cylinder, if one examines the Seiberg–Witten equations on Xusing a family of metrics for which the size of the fibers shrinks to zero, then one obtains in the adiabatic limit the equations for a holomorphic family of solutions to the symplectic vortex equations on the fibers. In turn, there is a natural isomorphism between the space of solutions to the vortex equations on a Riemann surface and the symmetric product of the surface. In other words, in this simple context the adiabatic limit of the Seiberg–Witten equations is the equation for a holomorphic family of elements of the symmetric products of the fibers of the fibrationX→ R×S1. As was noted in [DS], since for a Lefschetz fibration f: X→ S2 DSf precisely counts pseudoholomorphic families of elements of the symmetric products of the fibers of f, one might take inspiration from Salamon’s example and hope to obtain the equivalence between DSf and the Seiberg–Witten invariant by considering the Seiberg–Witten equations onX for a family of metrics with respect to which the size of the fibers shrinks to zero.

Now our invariant FDSnf is constructed by counting pseudoholo- morphic families of elements of the symmetric products of the fibers of a family of Lefschetz fibrations fb obtained by restricting a map fn: Xn+1→ S2×Xn to the preimage Xb ofS2× {b} asb ranges over the complementXn of a set of codimension 4 inXn. In the above vein, one might hope to relate the family Seiberg–Witten invariants F SW for the family of 4-manifolds Xn+1 → Xn (which enumerate Seiberg–

Witten monopoles in the various Xb as b ranges over Xn; see, e.g., [LL]) to FDSnf via an adiabatic limit argument. This would in partic- ular yield a proof of the independence of FDSnf from f in Conjecture 1.3, and indeed may well be the most promising way to establish this independence in the absence of a suitable invariantGrn (or of a “family Gromov–Taubes invariant” F Gr) with whichFDSnf might be directly equated.

(9)

As was shown in [L2], whenX is an algebraic surface andb+(X) = 1 the family Seiberg–Witten invariants agree with certain curve counts in algebraic geometry. For larger values ofb+, though, the family Seiberg–

Witten invariants that are hoped to correspond to nodal curve counts are expected to vanish due to the fact that symplectic manifolds have Seiberg–Witten simple type; note that Theorem 1.5 suggests thatFDSnf also tends to vanish for large b+. By contrast, there are plenty of non- trivial nodal curve counts in algebraic surfaces withb+>1 (see [L1] for a review of some of these); these counts correspond to Liu’s “algebraic Seiberg–Witten invariants” ASW and differ from F SW when b+>1.

Acknowledgements.Section 2 of this paper appeared in my thesis [U2]; I would like to thank my advisor Gang Tian for suggesting that I attempt to study nodal curves using the Donaldson–Smith approach.

Thanks also to Cliff Taubes for helping me identify an error an an earlier version of this paper, to Dusa McDuff for making me aware of the need for Section 6.1, and to Ivan Smith for helpful remarks. This work was partially supported by the National Science Foundation.

2. Refining the standard surface count

Throughout this section,Xr(f) will denote the relative Hilbert scheme constructed from some high-degree but fixed Lefschetz fibrationf: X → S2 obtained by Donaldson’s construction applied to the fixed symplec- tic 4-manifold (X, ω). The fiber of f over t ∈ S2 will occasionally be denoted by Σt, and the homology class of the fiber by [Φ].

As has been mentioned earlier, DSf(α) is a count of holomorphic sections of the relative Hilbert scheme Xr(f) in a certain homotopy class cα characterized by the property that if sis a section in the class cα then the closed setCs⊂X“swept out” bys(that is, the union over all tof the divisorss(t)∈Σt) is Poincar´e dual toα(note that points of Csin this interpretation may have multiplicity greater than 1). Thatcα is the unique homotopy class with this property is seen in Lemma 4.1 of [Sm2]; in particular, for instance, we note that sections which descend toconnected standard surfaces Poincar´e dual toαare not distinguished at the level of homotopy from those which descend to disjoint unions of several standard surfaces which combine to representP D(α).

Of course, in studying standard surfaces it is natural to wish to know their connected component decompositions, so we will presently attempt to shed light on this. Suppose that we have a decomposition

α=α1+· · ·+αn

with

hα,[Φ]i=r, hαi,[Φ]i=ri.

(10)

Over each t∈S2 we have an obvious “divisor addition map”

+ : Yn i=1

SriΣt→SrΣt

(D1, . . . , Dn)7→D1+· · ·+Dn; allowingt to vary we obtain from this a map on sections:

+ : Yn i=1

Γ(Xri(f))→Γ(Xr(f)) (s1, . . . , sn)7→

Xn i=1

si. As should be clear, one has

+(cα1 × · · · ×cαn)⊂cα if α=P

αi, since CPαi is the union of the standard surfaces Csi and hence is Poincar´e dual to α if eachCsi is Poincar´e dual to αi. Further, we readily observe:

Lemma 2.1. The image+(cα1×· · ·×cαn)⊂cα is closed with respect to the C0 norm.

Proof. Suppose we have a sequence (sm1 , . . . , smn)m=1incα1× · · · ×cαn such that P

smi → s ∈ cα. Now each SriΣt is compact, so at each t, each of the sequencessmi (t) must have subsequences converging to some s0i(t). But then necessarily each P

s0i(t) = s(t), and then we can see by, for anyl, fixing the subsequence used for alli6=l and varying that used for i=l that in fact every subsequence ofsml (t) must converge to s0l(t). Lettingtvary then gives sectionss0i such that everysmi →s0i and Ps0i =s; the continuity of sis readily seen to imply that of the s0i. q.e.d.

At this point it is useful to record an elementary fact about the lin- earization of the divisor addition map.

Proposition 2.2. Let Σ be a Riemann surface and r = P

ri. The linearization + of the addition map

+ : Yn i=1

SriΣ→SrΣ

at(D1, . . . , Dn)is an isomorphism if and only ifDi∩Dj =∅for i6=j.

If two or more of the Di have a point in common, then the image of+ at(D1, . . . , Dn) is contained in TPDi∆, where∆⊂SrΣis the diagonal stratum consisting of divisors with a repeated point.

(11)

Proof. By factoring + as a composition

Sr1Σ×Sr2Σ× · · · ×SrnΣ→Sr1+r2Σ× · · · ×SrnΣ→ · · · →SrΣ in the obvious way we reduce to the case n= 2. Now in general for a divisor D = P

aipi ∈ SdΣ where the pi are distinct, a chart for SdΣ is given by Q

SaiUi, where the Ui are holomorphic coordinate charts around pi and the SaiUi use as coordinates the elementary symmetric polynomialsσ1, . . . , σai in the coordinates ofUiai. As such, ifD1andD2 are disjoint, a chart around D1+D2∈Sr1+r2Σ is simply the Cartesian product of charts around D1 ∈ Sr1Σ and D2 ∈ Sr2Σ, and the map + takes the latter diffeomorphically (indeed, biholomorphically) onto the former, so that (+)(D1,D2) is an isomorphism.

On the other hand, note that

+ : SaC×SbC→Sa+bC

is given in terms of the local elementary symmetric polynomial coordi- nates around the origin by

1, . . . , σa, τ1, . . . , τb)7→(σ11, σ21τ12, . . . , σaτb), and so has linearization

(+)1,...,τb)1, . . . , ηa, ζ1, . . . , ζb)

= (η11, η21ζ11η12, . . . , σaζbbηa).

We thus see that Im(+)(0,...,0) only has dimension max{a, b} and is contained in the image of the linearization of the smooth model

C×Sa+b−2C→Sa+bC (z, D)7→2z+D

for the diagonal stratum at (0,0+· · ·+0). Suppose now thatD1andD2

contain a common pointp; writeDi =aip+Di whereDi∈Sri−aiΣ are divisors which do not containp. Then from the commutative diagram

Sa1Σ×Sr1−a1Σ×Sa2Σ×Sr2−a2Σ −−−−→ Sr1Σ×Sr2Σ

 y

 y+ Sa1+a2Σ×Sr1+r2−a1−a2Σ −−−−→ Sr1+r2Σ

and the fact that the linearization of the top arrow at (a1p, D1, a2p, D2) is an isomorphism (by what we showed earlier, since the Di do not contain p), while the linearization of the composition of the left and bottom arrows at (a1p, D1, a2p, D2) has image contained in TD1+D2∆, it follows that (+)(D1,D2) has image contained in TD1+D2∆ as well, which suffices to prove the proposition. q.e.d.

Corollary 2.3. Ifsi ∈Γ(Xri(f))are differentiable sections such that Csi∩Csj 6=∅ for some i6=j, then s=P

si ∈Γ(Xr(f)) is tangent to the diagonal stratum of Xr(f).

(12)

Proof. Indeed, if Csi∩Csj 6= ∅, then there is x ∈ S2 such that the divisorssi(x) andsj(x) contain a point in common, and so forv∈TxS2 we have

sv= (+◦(si, sj))v= +(s1∗v, s2∗v)∈Ts(t)

by Proposition 2.2. q.e.d.

Note that it is straightforward to find cases in which thesi are only continuous with some Csi ∩Csj nonempty and the sum s = P

si is smooth but not tangent to the diagonal. For example, letr = 2, and in local coordinates let s1 be a square root of the functionz7→Re(z) and s2 =−s1. Then in the standard coordinates on the symmetric product we have s(z) = (0,−Re(z)), so thatT(Ims) shares only one dimension with T∆ at z= 0. If s istransverse to ∆, one can easily check that a similar situation cannot arise.

We now bring pseudoholomorphicity in the picture. Throughout this treatment, all almost complex structures on Xr(f) will be assumed to agree with the standard structures on the symmetric product fibers, to make the map F: Xr(f) → S2 pseudoholomorphic, and, on some (not fixed) neighborhood of the critical fibers of F, to agree with the holomorphic model for the relative Hilbert scheme over a disc around a critical value for f provided in Section 3 of [Sm2]. Let J denote the space of these almost complex structures. It follows by standard argu- ments (see Proposition 3.4.1 of [MS1] for the general scheme of these arguments and Section 4 of [DS] for their application in the present context) that for genericJ ∈ J the spaceMJ(cα) is a smooth manifold of (real) dimension 2d(α) = α2−κX ·α (the dimension computation comprises Lemma 4.3 of [Sm2]); this manifold is compact, for bubbling is precluded by the arguments of Section 4 of [Sm2] assuming we have taken a sufficiently high-degree Lefschetz fibration.

InsideMJ(cα) we have the setMJ(cα1×· · ·×cαn) consisting of holo- morphic sections which lie in the image +(cα1× · · · ×cαn). By Lemma 2.1 and the compactness of MJ(cα), MJ(cα1 × · · · ×cαn) is evidently compact; however, the question of its dimension or even whether it is a manifold appears to be a more subtle issue in general.

Let us pause to consider what we would like the dimension ofMJ(cα1×

· · · × cαn) to be. The objects in MJ(cα1 × · · · ×cαn) are expected to correspond in some way to unions of holomorphic curves Poincar´e dual to αi. Accordingly, assume we have chosen theαi so that d(αi) =

1

22i−κX·αi)≥0 (for otherwise we would expectMJ(cα1×· · ·×cαn) to be empty). Holomorphic curves in these classes will intersect positively as long as they do not share any components of negative square; for a generic almost complex structure the only such components that can

(13)

arise are (−1)-spheres, so if we choose the αi to not share any (−1)- sphere components (i.e., if the αi are chosen so that there is no classE represented by a symplectic (−1)-sphere such thathαi, Ei<0 for more than one αi), then it would also be sensible to assume that αi·αj ≥0 fori6=j.

The above naive interpretation ofMJ(cα1× · · · ×cαn) would suggest that its dimension ought to be P

d(αi). Note that d(α) =d(X

αi) =X

d(αi) +X

i>j

αi·αj,

so under the assumptions on theαifrom the last paragraph we have that the expected dimension of MJ(cα1 × · · · ×cαn) is at most the actual dimension of MJ(cα) (as we would hope, given that the former is a subset of the latter), with equality if and only if αi·αj = 0 whenever i6=j.

As usual, we will find it convenient to cut down the dimensions of our moduli spaces by imposing incidence conditions, so we shall fix a set Ω of pointsz∈Xand consider the spaceMJ,Ω(cα1×· · ·×cαn) of elements s∈ MJ(cα1× · · · ×cαn) such thatCs passes through each of the points z (or, working more explicitly inXr(f), such thats meets each divisor z+Sr−1Σt, Σt being the fiber which contains z). MJ,Ω(cα) is defined similarly, and standard arguments show that for generic choices of Ω MJ,Ω(cα) will be a compact manifold of dimension

2(d(α)−#Ω).

We wish to count J-holomorphic sections s of Xr(f) such that the reducible components of Cs are Poincar´e dual to the αi. If we impose Pd(αi) incidence conditions, then according to the above discussion MJ,Ω(cα) will be a smooth manifold of dimension 2P

i>jαi ·αj. A section P

si ∈ +(cα1 × · · ·cαn) whose summands are all differentiable would then, by Corollary 2.3, have one tangency to the diagonal ∆ for each of the intersections between the Csi, of which the total expected number is P

i>jαi ·αj. This suggests that the sections we wish to count should be found among those elements of MJ,Ω(cα) which have P

i>jαi·αj tangencies to ∆, where Ω is a set of P

d(αi) points.

To count pseudoholomorphic curves tangent to a symplectic subvari- ety it is necessary to restrict to almost complex structures which pre- serve the tangent space to the subvariety (see [IP2] for the general the- ory when the subvariety is a submanifold). Accordingly, we shall restrict attention to the class of almost complex structures J on Xr(f) which arecompatible with the stratain the sense to be explained presently (for more details, see Section 6 of [DS], in which the notion was introduced).

(14)

Within ∆, there are various strata χπ indexed by partitions π :r = Paini with at least oneai >1; these strata are the images of the maps

pχ: Xn1(f)×S2· · · ×S2 Xnk(f)→Xr(f) (D1, . . . , Dk)7→X

aiDi;

in particular, ∆ = χr=2·1+1·(r−2). An almost complex structure J on Xr(f) is said to be compatible with the strata if the mapspχare (J, J)- holomorphic for suitable almost complex structuresJ on their domains.

Denoting by Yχ the domain of pχ, Lemma 7.4 of [DS] and the dis- cussion preceding it show:

Lemma 2.4([DS]). For almost complex structuresJ onXr(f)which are compatible with the strata, each J-holomorphic section s of Xr(f) lies in some unique minimal stratum χ and meets all strata contained in χ in isolated points. In this case, there is a J-holomorphic sec- tion s of Yχ such that s=pχ◦s. Furthermore, for generic J among those compatible with the strata, the actual dimension of the space of all such sections s is equal to the expected dimension of the space of J-holomorphic sectionss lying over s.

We note the following analogue for standard surfaces of the positivity of intersections of pseudoholomorphic curves.

Proposition 2.5. Let s = m1s1 +· · ·+mksk be a J-holomorphic section ofXr(f), where thesi ∈cαi ⊂Γ(Xri(f))are each not contained in the diagonal stratum of Xri(f), and where the almost complex struc- ture J on Xr(f) is compatible with the strata. Assume that the si are all differentiable. Then all isolated intersection points of Csi and Csj contribute positively to the intersection number αi·αj.

Proof. We shall prove the lemma for the casek= 2, the general case being only notationally more complicated. The analysis is somewhat easier if the points of Cs1 ∩Cs2 ⊂ X at issue only lie over t ∈ S2 for which s1(t) and s2(t) both miss the diagonal of Xr1(f) and Xr2(f), respectively, so we first argue that we can reduce to this case. Let χ be the minimal stratum (possibly all of Xr(f)) in which s = m1s1 + m2s2 is contained, so that all intersections of s with lower strata are isolated. Let p ∈ X be an isolated intersection point of Cs1 and Cs2 lying over 0 ∈ S2, and let δ > 0 be small enough that there are no other intersections of swith any substrata ofχ(and so in particular no other points ofCs1∩Cs2) lying overD(0)⊂S2. We may then perturb s=m1s1+m2s2 to ˜s=m11+m22, still lying inχ, such that

(i) Over Dδ(0), ˜s is J-holomorphic and disjoint from all substrata having real codimension larger than 2 inχ, and the divisors ˜s1(0) and ˜s2(0) both still containp;

(ii) Over the complement ofD(0), ˜sagrees with s; and

(15)

(iii) Over D(0)\ Dδ(0), ˜s need not be J-holomorphic but is con- nected to s by a family of sections st contained inχ which miss all substrata of χ

(it may be necessary to decrease δ to find such ˜s, but after doing so such ˜swill exist by virtue of the abundance of J-holomorphic sections over the small disc Dδ(0) which are close to s|Dδ(0)). The contribution of p to the intersection number α1·α2 will then be equal to the total contribution of all the intersections of Cs˜1 and Cs˜2 lying over Dδ(0), and the fact that ˜smisses all substrata with codimension larger than 2 inχis easily seen to imply that these intersections (of which there is at least one, at p) are all at points where ˜s1 and ˜s2 miss the diagonals in Xr1(f) andXr2(f).

As such, it suffices to prove the lemma for intersection points at which s1 and s2 both miss the diagonal. In this case, in a coordinate neighborhood U aroundp, theCsi can be written as graphsCsi ∩U = {w = gi(z)}, where w is the holomorphic coordinate on the fibers of X, z is the pullback of the holomorphic coordinate on S2, and gi is a differentiable complex-valued function which vanishes at z = 0.

Suppose first that m1 = m2 = 1. Then near s(0), we may use co- ordinates (z, σ1, σ2, y3, . . . , yr) for Xr(f) obtained from the splitting T0S2 ⊕ T2pS2Σ0 ⊕Ts(t)−2pSr−2Σ0, and the first two vertical coordi- nates of s(z) = (s1 +s2)(z) with respect to this splitting are (g1(z) + g2(z), g1(z)g2(z)). NowsisJ-holomorphic and meets theJ-holomorphic diagonal stratum ∆ at (0, s(0)), and at this point ∆ is tangent to the hy- perplane σ2 = 0, so it follows from Lemma 3.4 of [IP2] that the Taylor expansion of g1(z)g2(z) has form a0zd+O(d+ 1). But then the Tay- lor expansions of g1(z) and g2(z) begin, respectively, a1zd1+O(d1+ 1) and a2zd2 +O(d2+ 1), withd1+d2 =d. Then since Csi ∩U ={w = gi(z)}, it follows immediately that theCsi have intersection multiplicity max{d1, d2}>0 at p.

There remains the case where one or both of themi is larger than 1.

In this case, where Yχ =Xr1(f)×S2 Xr2(f) is the smooth model for χ, becauseJ is compatible with the strata, (s1, s2) is aJ-holomorphic sec- tion ofYχfor an almost complex structureJsuch thatpχ: Yχ→Xr(f) is (J, J)-holomorphic. Now where ˜∆ ={(D1, D2)∈Yχ|D1∩D2 6=∅}, compatibility with the strata implies that ˜∆ will beJ-holomorphic. In a neighborhoodV around (s1(z), s2(z)), we have, in appropriate coordi- nates, ˜∆∩V ={(z, w, w, D1, D2)|w∈Σz},while (s1(z), s2(z)) has first three coordinates (z, g1(z), g2(z)). From this it follows by Lemma 3.4 of [IP2] that

g1(z)−g2(z) =a0zd+O(d+ 1)

for some d, in which case Cs1 and Cs2 have intersection multiplicity

d >0 atp. q.e.d.

(16)

Definition 2.6. Let Ω be a set of P

d(αi) points and let J be an almost complex structure compatible with the strata. MJ,Ω01, . . . , αn) shall denote the set of J-holomorphic sections s ∈ cα with Ω ⊂ Cs such that there exist C1 sections si ∈ cαi with s = P

si, while the si themselves do not admit nontrivial decompositions as sums of C1 sections.

We would like to assert that for genericJ and Ω, the spaceMJ,Ω01, . . . , αn) does not include any sections contained within the strata. This is not true in full generality; rather we need the following assumption in order to rule out the effects of multiple covers of square-zero tori and (−1)-spheres in X.

Assumption 2.7. None of theαi can be written as αi =mβ where m >1 and either β2X·β= 0 orβ2X ·β=−1.

Under this assumption, we note that ifs=P

si ∈ MJ,Ω01, . . . , αn) were contained in ∆, then since the αi and hence the si are distinct we can write each si as si = mii with at least one mi > 1. The minimal stratum ofswill then beχπ whereπ =n

r=P mi³

ri

mi

´oand s = ( ˜s1, . . . ,s˜n) will be a J-holomorphic section ofYχ withs=pχ◦s, in the homotopy class [cα1/m1× · · · ×cαn/mn].

If any of thed(αi/mi)<0, then Lemma 2.4 implies that there will be no such sections s at all; otherwise (again by Lemma 2.4) the real di- mension of the space of such sections (taking into account the incidence conditions) will be

(2.1) 2³X

d(αi/mi)−X d(αi

.

But an easy manipulation of the general formula for d(β) and the ad- junction formula (which applies here because the standard surface cor- responding to a section of Xr(f) which meets ∆ positively will be sym- plectic; c.f. Lemma 2.8 of [DS]) shows that if d(β)≥0 andm≥2 then d(mβ)> d(β) unless either β2X ·β = 0 or β2X ·β =−1, and these are ruled out in this context by (i) and (ii) above, respectively. So Assumption 2.7 implies that the dimension in Equation 2.1 is negative, so no such s will exist for genericJ. This proves part of the following:

Proposition 2.8. Under Assumption 2.7, for generic pairs (J,Ω) where J is compatible with the strata and#Ω =P

d(αi),MJ,Ω01, . . . , αn) is a finite set consisting only of sections not contained in ∆.

Proof. That no member ofMJ,Ω01, . . . , αn) is contained in ∆ follows from the above discussion. As for the dimension of our moduli space, note that any s=P

si ∈ MJ,Ω01, . . . , αn) has one tangency (counted with multiplicity) to ∆ for each of the intersections of the Csi, of which there are P

αi ·αj (counted with multiplicity; this multiplicity will

(17)

always be positive by Proposition 2.5). By the results of Section 6 of [IP2], the space MJ,Ωδ,∆(cα) of J-holomorphic sections in the class cα having δ tangencies to ∆ and whose descendant surfaces pass through Ω will, for generic (J,Ω), be a manifold of dimension

2(d(α)−X

d(αi)−δ) = 2³X

αi·αj −δ´ , which is equal to zero in the case δ =P

αi·αj of present relevance to us.

Let us now show thatMJ,Ω01, ..., αn) is compact. Now since +(cα1×

· · · ×cαn) is C0-closed in cα, by Gromov compactness any sequence s(m) = Pn

i=1s(m)i in MJ,Ω01, . . . , αn) has (after passing to a subse- quence) a J-holomorphic limit s=P

si where the si∈cαi are at least continuous. We claim that, at least for generic (J,Ω), we can guarantee the si to be C1. In light of Proposition 2.2, the differentiability of the si is obvious at all points wheresmisses the diagonal, sincesis smooth by elliptic regularity and the divisor addition map induces an isomor- phism on the tangent spaces away from the diagonal. Now each s(m) hasP

αi·αj tangencies to the diagonal, corresponding to pointst∈S2 at which some pair of the divisors s(m)i (t) share a point in common.

The limit s will then likewise have n tangencies to the diagonal; the dimension formulas in [IP2] ensure that for generic (J,Ω) no two of the tangencies will coalesce into a higher order tangency to the smooth part of ∆ in the limit, and all of the intersecions ofIm swith the smooth part of the diagonal other than these n tangencies will be transverse. Fur- thermore, one may easily show (using for instance an argument similar to the one used in Lemma 2.1 of [U1] to preclude generic 0-dimensional moduli spaces of pseudoholomorphic curves in a Lefschetz fibration from meeting the critical points) that since the singular locus of ∆ has codi- mension 4 inXr(f), ifJhas been chosen generically thenswill not meet

sing, and so no s(t) will contain more than one repeated point (and that point cannot appear with multiplicity larger than two). In light of this, each tangency of s to ∆ will occur at a point s(t) where some pair si(t) and sj(t) have some point p in common, and all other points contained in anysk(t) are distinct from each other and fromp. Thanks to Proposition 2.2, this effectively reduces us to the case r = 2, with s=s1+s2 a sum of continuous sections with s1(0) =s2(0) = 0 which is holomorphic with respect to an almost complex structure which pre- serves the diagonal stratum ∆ inD2×Sym2D2, such thatsis tangent to ∆. Then lettingδ(t) = (s1+s2)2(t)−4s1(z)s2(t) be the discriminant, that s is tangent to the diagonal stratum implies, using Lemma 3.4 of [IP2], that δ(t) = at2 +O(3) for some constant a; in particular δ(t) has twoC1 square roots±r(t). Sincesis smooth, so is its first coordi- natet7→s1(t) +s2(t); adding this smooth function to the C1 functions

Références

Documents relatifs

As we will see in § 4, the symplectic structures corresponding to anti- self-dual metrics with negative definite Ricci operator fail to be K¨ ahler for precisely the same reason as

From the construction of VFC for M 0 1,k (X, A; J ) in Subsection 1.4, it is imme- diate that the difference between the standard and reduced genus-one GW-invariants of X must be

The Kumasi Metropolitan Assembly's Waste Management Department (KMA WMD) tried to promote private household toilets in the city of Kumasi according to national policy, but

The IFT gives a rather clear picture of the local structure of the moduli space of minimal immersions. In this section we want to use some notions of the theory of

the only reducible curves which satisfy one of these two conditions have two irreducible components, both real, and only transversal double points as sin- gularities with the

Abstract Let |L | be a linear system on a smooth complex Enriques surface S whose general member is a smooth and irreducible curve of genus p, with L 2 &gt; 0, and let.. V |L|,δ (S)

Magma codes for ”Some singular curves and surfaces arising from invariants of complex reflection groups”... “SOME SINGULAR CURVES AND SURFACES ARISING FROM INVARIANTS OF

Then the closure of a general leaf of the characteristic foliation on Y is a fiber of π (hence an abelian variety of dimension n).. In order to prove Theorem 0.4 we are going to